Vinylic C−H Bond Activation and Hydrogenation Reactions of Tp'Ir

Robert G. Bergman, Thomas R. Cundari, Aaron M. Gillespie, T. Brent Gunnoe, W. Dean Harman, Thomas R. Klinckman, M. Danielle Temple, and David P. White...
1 downloads 0 Views 179KB Size
4538

Inorg. Chem. 1998, 37, 4538-4546

Vinylic C-H Bond Activation and Hydrogenation Reactions of Tp′Ir(C2H4)(L) Complexes Enrique Gutie´ rrez-Puebla,† A Ä ngeles Monge,† M. Carmen Nicasio,‡ Pedro J. Pe´ rez,‡ ,§ Manuel L. Poveda,* Luis Rey,§ Caridad Ruı´z,† and Ernesto Carmona*,§ Instituto de Ciencia de Materiales de Madrid, Consejo Superior de Investigaciones Cientı´ficas, Campus de Cantoblanco, 28049 Madrid, Spain, Departamento de Quı´mica y Ciencia de los Materiales, Universidad de Huelva, 21819 Palos de la Frontera, Huelva, Spain, and Departamento de Quı´mica Inorga´nicasInstituto de Investigaciones Quı´micas, Universidad de Sevilla-Consejo Superior de Investigaciones Cientı´ficas, c/ Ame´rico Vespuccio s/n, Isla de la Cartuja, 41092 Sevilla, Spain ReceiVed January 22, 1998 The substitution of one of the ethylene ligands of the complexes Tp′Ir(C2H4)2 (Tp′ ) TpMe2, 1*; Tp′ ) Tp, 1) by soft donors such as tertiary phosphines or carbon monoxide is a facile reaction that gives the corresponding Tp′Ir(C2H4)(L) adducts. Spectroscopic studies support their formulation as five-coordinate, 18-electron species that possess a distorted trigonal bipyramidal geometry. This proposal has been confirmed by a single-crystal X-ray study carried out with the PMe2Ph complex TpMe2Ir(C2H4)(PMe2Ph) (3b*). Related hydride derivatives of Ir(III) can be obtained either by hydrogenation of the Ir(I) adducts (in general, this gives Tp′IrH2(L) compounds) or by thermal activation of one of the C-H bonds of the coordinated C2H4 ligand of the TpMe2Ir(C2H4)(L) compounds. All these reactions can be understood by invoking the participation of transient, 16-electron (η2Tp′)Ir intermediates, but the thermodynamics of the [Ir](C2H4) to [Ir]H(CHdCH2) conversion does not require an overall change in the coordination mode of the Tp′ ligand.

Introduction The transformation of transition metal-ethylene complexes into their hydride-vinyl isomers is, in general, thermodynamically uphill for mononuclear systems.1-3 Up to now, the only exceptions to this rule which are known involve Tp′Ir systems (Tp′ ) hydrotris(1-pyrazolyl)borate ligand4). Thus, the Ir(III) hydride-vinyl derivative TpCF3,MeIrH(CHdCH2)(CO)5 and the somewhat analogous TpMe2IrH(CHdCH2)(C2H4) complex and others closely related to it6 were found to be the products of †

Instituto de Ciencia de Materiales de Madrid, CSIC. Universidad de Huelva. Universidad de Sevilla-CSIC. (1) (a) Bell, T. W.; Haddleton, D. M.; McCamley, A.; Partridge, M. G.; Perutz, R. N.; Willner, H. J. Am. Chem. Soc. 1990, 112, 9212. (b) Bell, T. W.; Brough S.-A.; Partridge, M. G.; Perutz, R. N.; Rooney, A. D. Organometallics 1993, 12, 2933. (c) Bianchini, C.; Barbaro, P.; Meli, A.; Peruzzini, M.; Vacca, A.; Vizza, F. Organometallics 1993, 12, 2505. (2) Hydride-vinyl species however can be active in the evolution of M-ethylene complexes. See for example: (a) Desrosiers, P. J.; Cai, L.; Halpern, J. J. Am. Chem. Soc. 1989, 111, 8513. (b) Burger, P.; Bergman R. G. J. Am. Chem. Soc. 1993, 115, 10462. (c) Pe´rez, P. J.; Poveda, M. L.; Carmona, E. Angew. Chem., Int. Ed. Engl. 1995, 34, 231. (3) For some recent reviews on C-H bond activation see: (a) Ryabov, A. D. Chem. ReV. 1990, 90, 403. (b) Arndsten, B. A.; Bergman, R. G.; Mobley, T. A.; Peterson, T. H. Acc. Chem. Res. 1995, 28, 154. (4) Abreviations used in this paper: Tp ) HB(pz)3, TpMe2 ) HB(3,5Me2pz)3, Tp′ ) any tris(pyrazolyl)borate ligand. For recent, general reviews on Tp′ ligands see: (a) Trofimenko, S. Chem ReV. 1993, 93, 943. (b) Parkin, G. AdV. Inorg. Chem. 1995, 42, 291. (c) Kitajima, N.; Tolman, W. B. Prog. Inorg. Chem. 1995, 43, 418. In these references, the values of the cone angles for the Tp and TpMe2 ligands are somewhat different. The numbers given in the text are taken from ref 4c. (5) Ghosh, C. K.; Hoyano, J. K.; Kreutz, R.; Graham, W. A. G. J. Am. Chem. Soc. 1989, 111, 5480. (6) Alvarado, Y.; Boutry, O.; Gutie´rrez, E.; Monge, A.; Nicasio, M. C.; Poveda, M. L.; Pe´rez, P. J.; Ruı´z, C.; Bianchini, C.; Carmona, E. Chem.sEur. J. 1997, 3, 860. ‡ §

the thermal activation of the corresponding Ir(I)-olefin species Tp′Ir(C2H4)(L). In Graham’s system,5 namely TpCF3,MeIr(C2H4)(CO),5 a change in the coordination mode of the Tp′ ligand from η2 to η3 was suggested to accompany the activation of the C-H bond of the C2H4 ligand, and it was further assumed that the extra coordination of the free pyrazolyl arm provided the thermodynamic driving force needed for the activation reaction to take place. We have demonstrated, however, that TpMe2Ir(C2H4)2 has a five-coordinate, 18-electron structure, both in solution and in the solid state.6 Hence, in our case, the above transformation occurs without a change in the net electron count at the Ir center and in the coordination mode of the TpMe2 ligand. In this contribution, we extend previous studies on olefinic C-H activation to a series of Tp′Ir(C2H4)(L) complexes (Tp′ ) Tp, TpMe2; L ) PR3, CO) and investigate in addition the hydrogenation of these compounds. Results and Discussion Synthesis and Characterization of Tp′Ir(C2H4)(L) Complexes. We showed recently that the compound TpMe2Ir(C2H4)2 (1*) reacts with hard donors such as MeCN and DMSO with formation of TpMe2Ir(CHdCH2)(C2H5)(L). These reactions require heating at 60 °C, with the participation of TpMe2IrH(CHdCH2)(C2H4) (2*) as an active intermediate.6 In marked contrast, soft bases such as tertiary phosphines and CO readily substitute one of the C2H4 ligands in 1* and give the corresponding TpMe2Ir(C2H4)(L) adducts, as illustrated in eq 1

S0020-1669(98)00078-0 CCC: $15.00 © 1998 American Chemical Society Published on Web 08/11/1998

Reactions of Tp′Ir(C2H4)(L) Complexes

Inorganic Chemistry, Vol. 37, No. 18, 1998 4539

for the monodentate phosphines PMe3 (3a*), PMe2Ph (3b*), and PEt3 (3c*). Interestingly, the chelating phosphine Me2PCH2CH2PMe2, dmpe, affords, under similar conditions, the binuclear species 3d*, in which the diphosphine ligand bridges the two equivalent metallic centers (eq 2).

Carbon monoxide also induces a fast substitution reaction in THF to give first the mixed adduct TpMe2Ir(C2H4)(CO) (4*) (NMR evidence) and then the hydride-metallocarboxylic compound 5* (eq 3), which has been fully characterized by IR and

NMR spectroscopy and microanalysis. We presume that complex 5* is formed by the action of adventitious water on the undetected dicarbonyl TpMe2Ir(CO)2, a compound previously reported in preliminary form.7 During the progress of our work, compound 5* was isolated by Venanzi et al.8 The analogous, albeit slower, reaction of the TpIr(CO)2 derivative with H2O has been investigated by Oro and co-workers.9 As stated by these authors, the very high reactivity of the Tp′Ir(CO)2 complexes toward H2O is remarkable.8,9 We will briefly come back later to this point. For the time being, it should be mentioned that, on a practical basis, the mixed C2H4-CO adduct 4* is best obtained by action of KTpMe2 on [IrCl(coe)(CO)]2 (coe ) cyclooctene) in the presence of C2H4. As expected, the bis(ethylene) derivative of the unsubstituted Tp ligand, TpIr(C2H4)2 (1),10 is also prone to undergo a similar C2H4-PR3 substitution, the analogous complexes 3b,c being formed readily according to eq 4. The similar PPh3 adduct has been obtained recently by Heinekey et al.11

All the compounds of composition Tp′Ir(C2H4)(L) exhibit spectroscopic properties in agreement with a common, trigonal bipyramidal geometry A. This structure is closely related to that found for 1 and 1*6 and results from the replacement of the axial C2H4 ligand of the latter compounds by the L group. Theoretical calculations by Eisenstein and Caulton show that in this way back-donation to the π* orbital of the remaining C2H4 group is maximized while keeping at a minimum the overall molecular electronic energy.12 This type of structure has also been suggested for TpIr(C2H4)(CO) and TpIr(C2H4)(7) Ball, R. G.; Ghosh, C. K.; Hoyano, J. K.; McMaster, A. D.; Graham, W. A. G. J. Chem. Soc., Chem. Commun. 1989, 341. (8) Venanzi, L. M. Personal communication. (9) Ferna´ndez, M. J.; Rodrı´guez, M. J.; Oro, L. A. J. Organomet. Chem. 1992, 438, 337. (10) (a) Ferna´ndez, M. J.; Rodrı´guez, M. J.; Oro, L. A.; Lahoz, F. J. J. Chem. Soc., Dalton Trans. 1989, 2073. (b) Tanke, R. S.; Crabtree, R. H. Inorg. Chem. 1989, 28, 3444. (11) Oldham, W. J.; Heinekey, D. M. Organometallics 1997, 16, 467.

(PPh3).11,13 As discussed below, the structure proposed for these complexes on spectroscopic grounds has been confirmed by an X-ray study carried out with the PMe2Ph complex 3b*. At variance with the parent bis(ethylene) compounds 1 and 1*, all the Tp′Ir(C2H4)(L) derivatives reported in this paper are rigid at room temperature on the NMR time scale. This is evidenced by the observation of two sets of resonances (2:1 intensity ratio) for the pyrazolyl groups of the Tp′ ligand. In fact, of the members of this Tp′Ir(C2H4)(PR3) family of compounds, only TpIr(C2H4)(PPh3) has been found to be fluxional in solution.11 The dynamic behavior of this complex could be ascribed to the steric pressure exerted by the bulky PPh3 ligand, which would facilitate the fast dissociation of one of the pyrazolyl arms. It is worth noting that even the coordinated C2H4 group of these mixed C2H4-L adducts has a static orientation, no rotation around the Ir-C2H4 axis taking place on the NMR time scale (an AA′BB′ spin pattern is observed for the ethylene protons under 31P-decoupling conditions). The mirror symmetry plane of these molecules is also manifested in the observation of only one resonance for the olefinic 13C nuclei. It is remarkable that this signal appears at rather high field, as compared with those of other related M-C2H4 complexes. For the TpMe2 series, the shielding increases with the donor capability of the L ligand: 0.6 (CO); -7.2 (PMe2Ph); -8.1 (PMe3); -9.1 (dmpe); -10.9 ppm (PEt3). These signals are observed at somewhat lower field for the analogous compounds of the less-donating, unsubstituted Tp group (e.g., 0.5 ppm in the TpIr(C2H4)(PMe2Ph) derivative) but seem to maintain the same dependence on the donicity of the L ligand (a chemical shift of 2 ppm has been reported for TpIr(C2H4)(PPh3)11). The one-bond 13C-1H coupling constant found for the C2H4 ligand in these complexes has a relatively low value of ca. 145 Hz. On the basis of this and of the above chemical shift data, it is tempting to speculate on the possibility that the metallacyclopropane resonance form B has an important contribution to the ground-state electronic structure of these compounds.14 Comparison with the data reported for other compounds that appear to behave chemically as metallacyclopropanes seems appropriate. For example, (C5H5)2Ti(C2H4),15a (ArO)2Ti(C2H4)(PMe3),15b and (C5H5)(Me2PCH2CR2O)Ti(C2H4)15c exhibit δ(C2H4) at 105, 72 (average), and 57 (average) ppm, respectively, and are further characterized by 1JCH values in the proximity of 145-150 Hz.15 The above δ(13C) values highlight the difficulties that arise when the attempt is made to compare the chemical shifts of a certain functionality bound to very (12) Lundquist, E. G.; Folting, K.; Streib, W. E.; Huffman, J. C.; Eisenstein, O.; Caulton, K. G. J. Am. Chem. Soc. 1990, 112, 858. (13) Ciriano, M. A.; Ferna´ndez, M. J.; Modrego, J.; Rodrı´guez, M. J.; Oro, L. A. J. Organomet. Chem. 1993, 443, 249. (14) Bender, B. R.; Norton, J. R.; Miller, M. M.; Anderson, O. P.; Rappe´, A. K. Organometallics 1992, 11, 3427. (15) (a) Cohen, S. A.; Auburn, P. R.; Bercaw, J. E. J. Am. Chem. Soc. 1983, 105, 1136. (b) Hill, J. E.; Fanwick, P. E.; Rothwell, I. P. Organometallics 1992, 11, 1171. (c) van Doorn, J. A.; van der Heijden, H.; Orpen, A. G. Organometallics 1995, 14, 1278. (d) For a discussion of 1JCC values see: Barry, J. T.; Chacon, S. T.; Chisholm, M. H.; Huffman, J. C.; Streib, W. E. J. Am. Chem. Soc. 1995, 117, 1974.

4540 Inorganic Chemistry, Vol. 37, No. 18, 1998

Gutie´rrez-Puebla et al.

Table 1. Crystal and Refinement Data for 3b* and 13b* 3b* formula cryst syst space group a, Å b, Å c, Å R, deg β, deg γ, deg V, Å3 Z F(000) Dcalcd, g cm-1 temp, °C µ(Mo KR), cm-1 cryst dimens, mm diffractometer radiation scan technique 2θ, range, deg data collcd no. of reflns collcd no. of unique data no. of obsd reflns Rint, % std reflns R1a wR2a av shift/error a

13b*

C25H37N6BPIr monoclinic C2/c 26.104(3) 16.767(5) 16.967(2) 128.551(9) 5808(2) 8 2608 1.5 22 46.6 0.4 × 0.2 × 0.2 Enraf-Nonius CAD4 graphite-monochromated Mo KR (λ ) 0.710 69 Å) ω/2θ (-31,0,0) to (31,19,20) 5269 5111 3210 (I g 2σI) 3.5 3/233 4.3 5.2 0.18

C25H37N6BIr triclinic P1h (N° 2) 11.287(3) 11.404(3) 12.516(3) 99.56(2) 100.73(2) 117.90(2) 1338.0(7) 1 652 1.63 22 50.5 0.1 × 0.2 × 0.3 Enraf-Nonius CAD4 graphite-monochromated Mo KR (λ ) 0.710 69 Å) ω/2θ 1-60 (-15,-16,0) to (15,16,17) 5674 3273 (I g 2σI) 5.8 3/92 4.7 5.1 0.06

R1 ) ∑||Fo| - |Fc||/∑|Fo|; wR2 ) [∑[w(Fo2 - Fc2)2/∑[w(Fo2)2]]1/2. Table 2. Selected Bond Lengths and Angles for 3b* Ir-P Ir-N12 Ir-N22 Ir-N32 Ir-C1 C1-Ir-C2 N32-Ir-C2 N32-Ir-C1 N22-Ir-C2 N22-Ir-C1 N22-Ir-N32 N12-Ir-C2 N12-Ir-C1 N12-Ir-N32 N12-Ir-N22 P-Ir-C2

Figure 1. Molecular structure of 3b*. Hydrogen atoms are omitted for clarity.

different metal environments. However, the coupling constants are very similar to those of our Ir compounds and may be considered in support of the metallacyclopropane formulation. Caution should nonetheless be taken, and we would rather stress that the chemical behavior of our Tp′Ir(C2H4)(L) complexes is that expected for Ir(I)-olefin species. It should also be mentioned that the somewhat related Cp′Ir(C2H4)(PMe3) compounds, which contain as well a static C2H4 ligand, are characterized by similar NMR parameters (Cp′ ) C5Me5, δ 10.5, 1J 16 Cp ) C H , δ 4.11a). We conclude that CH ) 151 Hz; 5 5

Bond Distances (Å) 2.207(4) Ir-C2 2.17(1) P-C3 2.16(1) P-C4 2.18(2) P-C5 2.10(2) C1-C2 Bond Angles (deg) 40.6(7) P-Ir-C1 153.6(5) P-Ir-N32 113.0(5) P-Ir-N22 112.9(7) P-Ir-N12 152.3(6) Ir-P-C5 93.2(5) Ir-P-C4 98.3 Ir-P-C3 94.6(5) C4-P-C5 81.6(5) C3-P-C5 79.7(4) C3-P-C4 90.0(6)

2.06(2) 1.82(2) 1.83(2) 1.82(1) 1.44(2) 93.3(4) 92.1(3) 94.7(3) 171.3(3) 117.8(4) 114.9(7) 119.8(5) 100.7(7) 101.6(8) 98.8(7)

retrodonation from the electron-rich Ir(I) center to the C2H4 ligand must be important in these compounds and suggest that this explanation should suffice to account for their NMR properties. As indicated above, a single-crystal X-ray study of TpMe2Ir(C2H4)(PMe2Ph) (3b*), has been undertaken. Figure 1 shows an ORTEP view of the molecules of this compound; a summary of the crystal data is given in Table 1, and pertinent bond distances and angles are summarized in Table 2. The Ir atom lies in the center of a severely distorted trigonal-bipyramidal (tbp) environment, similar to that found for the parent compound 1*, except, naturally, for the presence of a molecule of PMe2Ph in place of the axial ethylene group of the latter complex. As in 1*, the equatorial plane contains two of the N atoms of the TpMe2 group (which, as discussed below, form an almost (16) Stoutland, P. O.; Bergman, R. G. J. Am. Chem. Soc. 1988, 110, 5732.

Reactions of Tp′Ir(C2H4)(L) Complexes right N22-Ir-N32 angle of 93.2(5)°, far from the ideal value of 120° expected for a regular tbp geometry) as well as the carbon atoms C1 and C2 of the ethylene ligand. Within experimental error, the Ir-C1 and Ir-C2 separations are identical to the corresponding distances in 1* (2.08(2) Å (average) vs 2.06(2) Å in (1*)6). The C1-C2 bond length, 1.44(2) Å, is intermediate between a single (1.54 Å) and a double (1.34 Å) carbon-carbon bond. It is interesting to note that this bond has a length identical to that of the equatorial C2H4 group in 1*, which shows that the substitution of the axial C2H4 bond by the stronger donor PMe2Ph ligand has no apparent effect on this bonding parameter. A similar situation holds for the IrN12 bond, i.e. that trans to the PMe2Ph ligand, which is identical, within the limits of the standard deviation, to the other two (2.17(1), 2.16(1), and 2.18(1) Å). As mentioned above, the equatorial N-Ir-N angle deviates considerably from the ideal 120° value. This important distortion within the equatorial plane is also manifested by the very large N22-Ir-C2 and N32-Ir-C1 angles of 152.3 and 152.6°, respectively. The other two N-Ir-N bond angles also have values close to 90°, albeit slightly smaller (79.7(4) and 81.6(5)°), well in the range generally encountered in complexes that contain trihapto-bonded hydrotris(pyrazolyl)borate ligands.10a,17a We believe that this geometrical constraint imposed by the Tp′ ligands is largely responsible for the high tendency of these ligands to enforce six-coordination to the metal center.18 The structural data just discussed confirm the identity of the solution and solid-state structures of this compound and, by extension, of the other related compounds reported in this paper. Further confirmation comes from 13C{1H} CPMAS studies carried out with 3b* which show the C2H4 resonance at δ -8, i.e. very close to the solution value of -7.2 ppm. Before we conclude this section, some brief comments regarding the facility with which compounds 1 and 1* undergo substitution reactions in the presence of soft bases appear appropriate. Although both ethylene derivatives have a fivecoordinate, trigonal bipyramidal ground-state structure,6 fourcoordinate intermediates resulting from the disengagement of one of the pyrazolyl rings are probably sufficiently close in energy to become accessible at normal temperatures. These intermediates probably have a high affinity for the soft ligands whereas in the case of the harder donors (e.g. acetonitrile) an alternative reaction pathway, namely that involving vinylic C-H activation and formation of the hydride-vinyl species 2 and 2*, appears to be kinetically favored. The easy formation of the 16-electron intermediates is a characteristic of these Tp′IrI systems that makes them much more reactive, in particular in associative processes, than the corresponding Cp′IrI derivatives. This same conclusion has been independently reached by Heinekey and associates.11 Hydrogenation of Tp′Ir(C2H4)(L) Complexes. A characteristic chemical feature of the ethylene complexes of the Tp′IrI fragment is their ability to interact with H2 under very mild conditions. All the Tp′Ir(C2H4)(PR3) compounds tested react with H2 at 20 °C, under 1-2 atm of this gas, to yield quantitatively (by NMR monitoring) the new Ir(III) dihydrides (17) (a) Bovens, M.; Gerfin, T.; Gramlich, V.; Petter, W.; Venanzi, M. L.; Haward, M. T.; Jackson, S. A.; Eisenstein, O. New. J. Chem. 1992, 16, 337. (b) Ferrari, A.; Polo, E.; Ru¨egger, H.; Sostero, S.; Venanzi, L. M. Inorg. Chem. 1996, 35, 1602. (18) (a) Curtis, M. D.; Shiu, K. B.; Butler, W. M. J. Am. Chem. Soc. 1986, 108, 1550. (b) Curtis, M. D.; Shiu, K. B.; Butler, W. M.; Huffman, J. C. J. Am. Chem. Soc. 1986, 108, 3335. See also: Reger, D. L.; Huff, M. F.; Rheingold, A. L.; Haggerty, B. S. J. Am. Chem. Soc. 1992, 114, 579.

Inorganic Chemistry, Vol. 37, No. 18, 1998 4541 Tp′IrH2(PR3) (Tp′ ) TpMe2, PR3 ) PMe3 (6a*), PMe2Ph (6b*), 1/ dmpe (6d*); Tp′ ) Tp, PR ) PMe Ph (6b)). As shown in 2 3 2 eq 5, a related dihydride, 7* (L ) CO), is formed by starting

with the monocarbonyl complex 4*. The structure proposed for these derivatives is in agreement with their spectroscopic data, which are collected in the Experimental Section. Compound 6a* was obtained recently by Heinekey et al. using a different synthetic method.19 The dihydride complexes exhibit good thermal stability, but they decompose slowly in CDCl3. For example, solutions of compounds 6a* and 6b* in this solvent, when heated at 60-70 °C, convert slowly into the monohydrides TpMe2Ir(H)Cl(PR3) (PR3 ) PMe3 (8a*), PMe2Ph (8b*)) with concomitant production of CHDCl2. For synthetic purposes, it proves more convenient to heat the dihydrides in a mixture of CHCl3-CCl4 until complete transformation (NMR monitoring). A somewhat more complex behavior is found upon hydrogenation (1-2 atm) of the bis(ethylene) complex 1*, a mixture of two compounds being now formed (eq 6). One of them is

the dihydride species TpMe2IrH2(C2H4) (9*), related to those previously discussed and reported independently by Venanzi and co-workers.17 The second is the hydride ethyl compound 10*, whose formation is favored with respect to that of 9* upon lowering the temperature. Thus the 9*:10* ratio varies from 1:1.5 at 20 °C to 1:2.5 at 0 °C and 1:4 at -60 °C. These proportions are kinetic in origin since compound 10* does not react with H2 at the above temperatures (vide infra) and are in accord with the expected influence of the entropy term in the rates of the two competitive reactions, that involving the extrusion of one of the C2H4 ligands becoming more favorable at higher temperatures. At variance with a previous observation,10b the unsubstituted Tp complex 1 can also be hydrogenated at room temperature. In this case, however, only the ethyl complex TpIrH(C2H5)(C2H4) (10) appears to form. In accord with studies by Heinekey and co-workers,11 we propose that these hydrogenations proceed through an associative process in which 16-electron reactive intermediates are trapped by H2 to give species of type C, from which C2H4 may

be easily extruded. For reasons that remain to be fully understood, when L ) C2H4 the insertion of C2H4 into the Ir-H bond becomes kinetically competitive with the dissociation of the C2H4 ligand and ethyl products are obtained. (19) Oldham, W. J.; Hinkle, A. S.; Heinekey, D. M. J. Am. Chem. Soc. 1997, 119, 11028.

4542 Inorganic Chemistry, Vol. 37, No. 18, 1998

Gutie´rrez-Puebla et al.

Scheme 1

Compounds 9* and 10* are, in our opinion, interesting examples of Ir(III)-hydride-ethylene complexes that deserve further comment. As expected for an Ir(III)-ethylene linkage,1a,6 the coordinated olefin udergoes fast rotation on the NMR time scale around the Ir-C2H4 bond axis. The lower back-donation from the metal center, as compared to that of the analogous Ir(I) derivatives, is also manifested in the values of the chemical shift δ(C2H4) and of the coupling constants 1JCH (ca. 40 ppm and 160 Hz, respectively). The weakness of the Ir-C2H4 bond may also account for the facility with which insertion reactions take place. As shown in Scheme 1, treatment of 9* and 10* with PMe3, at 60 °C, produces the expected ethyl complexes 11* and 12*, respectively, as the result of the migratory insertion of ethylene into the Ir-H bond. In the case of the 10* to 12* conversion, this observation clearly implies that migration of the hydride group is more favorable than that of the ethyl fragment. Although this appears to be a general observation, not only for these TpMe2IrIII complexes6 but also for many organometallic compounds, we have found that the migratory insertion of C2H4 into an Ir-C bond can indeed take place under mild conditions.20 Scheme 1 also shows that at 60 °C the hydride-ethyl complex 10* reacts with H2 to produce the dihydride compound 9*. Further hydrogenation gives the known tetrahydride TpMe2IrH4.21 Interestingly, this transformation can be reversed; treatment of 9* with C2H4, at the same temperature, yields 10*. However, under these conditions, the latter compound undergoes formally a rapid σ-bond metathesis with C2H4 that gives first the hydride-vinyl complex TpMe2IrH(CHdCH2)(C2H4) (2*) and then, in a fast sequence of events, the products previously reported as resulting from the interaction of 2* with C2H4.6 Olefinic C-H Bond Activation in Tp′Ir(C2H4)(L) Complexes. As already mentioned, compounds 1 and 1* and other related species undergo thermal C-H vinylic activation, under mild conditions, to yield the corresponding hydride-alkenyl derivatives.6 We have now extended these studies to the related Tp′Ir(C2H4)(L) complexes and have found that in the TpMe2 series clean conversion to the hydride-vinyl species TpMe2IrH(CHdCH2)(L) takes place when L ) PMe3 or PMe2Ph, upon heating at 60-70 °C (C6D6, NMR monitoring). This conversion (eq 7) is about 1 order of magnitude faster for the PMe3 complex

as compared with the PMe2Ph analogue.22 For carbon monoxide, the transformation is somewhat disfavored; heating at (20) (a) Boutry, O.; Gutie´rrez, E.; Monge, A.; Nicasio, M. C.; Pe´rez, P. J.; Carmona, E. J. Am. Chem. Soc. 1992, 114, 7288. (b) Gutie´rrez-Puebla, E.; Monge, A.; Nicasio, M. C.; Pe´rez, P. J.; Poveda, M. L.; Carmona, E. Chem. Eur. J., in press. (21) (a) Paneque, M.; Poveda, M. L.; Taboada, S. J. Am. Chem. Soc. 1994, 116, 4519. (b) Paneque et al., submitted. (22) Surprisingly, these reactions are considerably slower when crude, noncrystallized, starting materials are used. We do not know the nature of the inhibitor, but a series of experiments indicates that it is not O2, H2O, PMe3, or adventitious acid or base.

Figure 2. Molecular structure of 13b*. Hydrogen atoms are omitted for clarity, except for H1 on iridium which is represented by a sphere of arbitrary radius. Table 3. Selected Bond Lengths and Angles for 13b* Ir-P Ir-C1 Ir-N12 Ir-N22 N22-Ir-N32 N12-Ir-N32 N12-Ir-N22 C1-Ir-N32 C1-Ir-N22 C1-Ir-N12 P-Ir-N32 P-Ir-N22

Bond Distances (Å) 2.242(3) Ir-N32 2.030(9) Ir-H1 2.105(9) C1-C2 2.201(9) Bond Angles (deg) 87.7(4) P-Ir-N12 86.2(4) P-Ir-C1 83.5(5) H1-Ir-N32 93.2(5) H1-Ir-N22 173.7(5) H1-Ir-N12 90.3(5) H1-Ir-C1 99.7(2) H1-Ir-P 96.7(3)

2.215(9) 1.6(2) 1.27(2)

174.1(3) 89.3(4) 163(6) 10(7) 103(6) 73(8) 71(6)

130 °C (cyclohexane) is needed for the reaction to proceed at a practical rate. The unsubstituted Tp ligand has also a negative effect; no clean product can be obtained when the complexes TpIr(C2H4)(PR3) (3b,c) are heated in C6D6 at 60-80 °C. An analogous behavior was also encountered for the parent TpIr(C2H4)2 derivative.6 We have not attempted the photochemical activation of 3b,c. In view of the scarcity of X-ray structures reported for hydride-vinyl complexes, we have carried a single-crystal X-ray analysis of the PMe2Ph derivative 13b*. This study appears further justified by the information it may provide on the intriguing reactivity of these alkenyl complexes of iridium, a topic which is under intense scrutiny in our laboratories.23 The structure (Figure 2) shows the expected distorted octahedral geometry, in which the N atoms of the TpMe2 ligand occupy three facial positions and the vinyl, hydride, and phosphine ligands occupy the others. Bond angles around the metal center (Table 3) have values close to those expected for octahedral geometry, except those involving the hydride ligand. However, due to the difficulties in locating this atom with sufficient precision, these deviations should be taken with caution. As in the parent compound 3b*, the N-Ir-N bond angles have values close to the ideal 90°; nevertheless, in 13b*, they are slightly acute (83.5(5), 86.2(4), and 87.7(4)°). The Ir-H distance of 1.6(2) Å is in the range reported for these bonds.6,10a,16,17,24 (23) (a) Alvarado, Y.; Daff, P. J.; Pe´rez, P. J.; Poveda, M. L.; Sa´nchezDelgado, R.; Carmona, E. Organometallics 1996, 15, 2192. (b) Alı´as, F. M.; Poveda, M. L.; Sellin, M.; Carmona, E. J. Am. Chem. Soc. 1998, 120, 5816. (c) Alı´as, F. M.; Poveda, M. L.; Sellin, M.; Carmona, E.; Gutie´rrez-Puebla, E.; Monge, A. Organometallics, in press.

Reactions of Tp′Ir(C2H4)(L) Complexes The Ir-P bond (2.242(3) Å) is longer than that in 3b* (2.207(4) Å). This is the opposite trend expected for the increase in the oxidation state of the Ir center, and whereas it might be attributed to a π-acceptor role25 of the PMe2Ph ligand in the Ir(I) complex 3b*, we would rather ascribe the lengthening of this bond in 13b* to its being trans with respect to the shortest Ir-N bonds (Ir-N12, 2.105(9) Å). As can be seen in Figure 2, the vinyl fragment lies almost in the plane defined by IrH1-N22-N32. It is actually slightly above that plane, almost eclipsing the H1-Ir-N32 vector, with CR facing the hydride ligand. This conformation is similar to that found in Cp*IrH(CHdCH2)(PMe3),16 but rotated by 180°. The Ir-C1 distance of 2.030(9) Å is however shorter than in the latter compound (2.054(4) Å), although this value can be considered normal. Finally, the length of the CdC bond in the vinyl group (1.270(2) Å) is somewhat shorter than that of a typical double bond but identical to that found in the Cp* complex mentioned above. It is pertinent to address again6 the C-H bond activation reaction that converts the Ir-C2H4 fragment into an isomeric IrH(CHdCH2) structure. The characterization of the TpMe2Ir(C2H4)(L) complexes as five-coordinate, 18-electron species leaves no doubt that the C-H activation occurs between electronically saturated, i.e. 18-electron, iridium centers. Hence, and at variance with Graham’s proposal for a somewhat related system,5 no thermodynamic driving force associated with the electronic and coordination unsaturation of the starting Ir(I) center can be invoked to explain the olefinic C-H activation reaction. In the related C5R5Ir-PMe3 system, the Ir-C2H4 and IrH(CHdCH2) structures exhibit opposite thermodynamics; i.e., the Ir-C2H4 complex is more stable than the hydride-vinyl isomer. Tp′ and Cp′ ligands have relatively similar electrondonor properties26a,b although some differences appear to exist and the following order of donor ability was recently proposed: 26c,d TpMe2 e C H < C Me . For the Ir(III) compounds (A)5 5 5 5 IrH(CHdCH2)(CO) (A ) TpMe2 (14*), C5H51a), almost identical ν(CO) frequencies have been identified (2020 and 2022 cm-1, respectively). A similar situation is encountered in Ir(I) compounds: 1990 cm-1 for TpMe2Ir(C2H4)(CO) (4*) and 1979 cm-1 for (C5H5)Ir(C2H4)(CO).1a The above data for the TpMe2 system show that an increase in ν(CO) of ca. 30 cm-1 accompanies the 4* to 14* transformation, and in this regard, it should be noted that a similar ∆ν(CO) of ca. 32 cm-1 is associated with the conversion of the TpCF3,MeIr(C2H4)(CO) (2030 cm-1) into TpCF3,MeIrH(CHdCH2)(CO) (2062 cm-1). This and the close resemblance of the 1H NMR data reported for TpCF3,MeIr(C2H4)(CO) with those of 4* may be taken as suggestive of analogous ground-state structures. Since, in TpIr(C2H4)(CO),13 ν(CO) appears at 2000 cm-1, the electrondonating power of these hydrotris(pyrazolyl)borate ligands varies in the order TpMe2 > Tp . TpCF3,Me.26e The Cp′ and Tp′ ligands differ considerably in size and therefore exert dissimilar steric pressure; cone angles of 236 (TpMe2), 199 (Tp), 182 (C5Me5), and 150° (C5H5) have been reported for these groups (Tp′;4c Cp′27). Even though these differences could be invoked to account for the above order of thermodynamic stability, i.e. [Ir]′H(CHdCH2) > [Ir]′(C2H4), (24) Fryzuk, M. D.; Gao, X.; Rettig, S. J. J. Am. Chem. Soc. 1995, 117, 3106. (25) Gilheany, D. G. Chem. ReV. 1994, 94, 1339. (26) (a) Sharp, P. R.; Bard, A. J. Inorg. Chem. 1983, 22, 2689. (b) Curtis, M. D.; Shiu, K. B. Inorg. Chem. 1985, 24, 1213. (c) Dunn, S. G.; Mountford, P.; Shishkin, O. V. Inorg. Chem. 1996, 35, 1006. (d) Koch, J. L.; Shapley, P. A. Organometallics 1997, 16, 4071. (e) Dias, H. V. R.; Lu, H.-L. Inorg. Chem. 1995, 34, 5380. (27) These values have been taken from ref 4e.

Inorganic Chemistry, Vol. 37, No. 18, 1998 4543 we would rather emphasize two additional factors which we believe play an important role in helping the Tp′Ir system to overcome the otherwise unfavorable thermodynamics of the Ir(C2H4) to IrH(CHdCH2) transformation: (a) the harder nature of the Tp′ ligands as compared to the Cp′, which makes them bind preferentially to the also harder Ir(III) centers, and (b) their well-known propensity to impose six-coordination at the metal center,18 a situation that is highly favorable for d6 Ir(III). We presume that these factors are of importance to understand the somewhat anomalous chemistry exhibited in this respect by the Tp′Ir complexes.6,17b,28 In the present case, these arguments can be additionally used to explain the exceedingly high reactivity of the dicarbonyl compounds Tp′Ir(CO)2 toward water. Since unsaturated species derived from the TpMe2Ir(C2H4)(L) compounds by dissociation of either the C2H4 or the L ligand are very active in reactions that involve aromatic C-H bond activation,29 such dissociation processes cannot take place during the vinylic C-H activations discussed above. A simple, concerted oxidative addition reaction would be in accord with all the experimental data accumulated during the progress of this work. However, a recent theoretical analysis by Hall et al.30 suggests the process could be more complex and require as a previous step the rupture of one of the Ir-N bonds within the TpMe2Ir fragment. A detailed study of these mechanistics aspects has not been undertaken, but nonetheless we have gathered enough qualitative evidence regarding the influence of the steric and electronic effects in the C-H bond activation reaction that may be in accord with this proposal. It appears reasonable to assume that an increase of the steric hindrance and/or the electron density at the metal center should favor the temporary disengagement of one of the pyrazolyl rings. When one compares the TpMe2 and the Tp systems, the bulkier and somewhat better donor TpMe2 favors the C-H activation,6 and the same can be said for the bulkier terminal olefins propene and 1-butene when they are compared with ethylene. Within the TpMe2 system, the reactivity increases in the order CO < PMe2Ph < PMe3 < C2H4, which clearly includes both electronic and steric effects. Finally, the latter seem predominant when a comparison of the reactivities of TpCF3,MeIr(C2H4)(CO)5 (the C-H activation occurs at 100 °C) and TpMe2Ir(C2H4)(CO) (4*) (120-140 °C) complexes is made. Conclusions Several complexes of the general composition Tp′Ir(C2H4)(L), for Tp′ ) Tp or TpMe2 and L ) tertiary phosphine or CO, have been isolated and characterized structurally as fivecoordinate, 18-electron species. Similar to other Tp′-C2H4 derivatives,5,6 but at Variance with the related Cp′ compounds, they are thermodynamically unstable with respect to their hydride-vinyl isomers. We propose that this behavior may be associated with the hard nature of the Tp′ ligands and with their strong tendency to impose six-coordination at the metal center, a situation which is particularly favorable for the d6 Ir(III) systems. The Ir(I) mixed C2H4-L adducts also undergo easy ligand exchange (with soft bases) and hydrogenation reactions, (28) (a) Gutie´rrez, E.; Monge, A.; Nicasio, M. C.; Poveda, M. L.; Carmona, E. J. Am. Chem. Soc. 1994, 116, 791. (b) Paneque, M.; Poveda, M. L.; Rey, L.; Taboada, S.; Carmona, E.; Ruı´z, C. J. Organomet. Chem. 1995, 504, 147. (c) Boutry, O.; Poveda, M. L.; Carmona, E. J. Organomet. Chem. 1997, 528, 143. (29) Unpublished results from this laboratory. (30) Jime´nez-Castan˜o, R.; Niu, S.; Hall, M. B. Organometallics 1997, 16, 1962. See also: Bromberg, S. E.; Yang, H.; Asplund, M. C.; Lian, T.; McNamara, B. K.; Kotz, K. T.; Yeston, J. S.; Wilkens, M.; Frei, H.; Bergman, R. G.; Harris, C. B. Science 1997, 278, 260.

4544 Inorganic Chemistry, Vol. 37, No. 18, 1998 which are best explained by the accessibility of low-energy, albeit undetected, 16-electron intermediates, formed by momentary unanchoring of one of the pyrazolyl arms. The same type of species may be the key intermediates in the vinylic C-H activation experienced by the TpMe2Ir(C2H4)(L) derivatives. Experimental Section Microanalyses were performed by the Analytical Service of the Universidad de Sevilla. Infrared spectra were obtained from PerkinElmer spectrometers, models 577 and 684. The NMR instruments were Varian XL-200, Bruker AMX-500, and Bruker AMX-300 spectrometers. Spectra were referenced to external SiMe4 (δ ) 0 ppm) using the residual protio solvent peaks as internal standards (1 H NMR experiments) or the characteristic resonances of the solvent nuclei (13C NMR experiments). Spectral assignments were made by means of routine one- and two-dimensional NMR experiments where appropriate. All manipulations were performed under dry, oxygen-free dinitrogen by following conventional Schlenk techniques. The complexes TpMe2Ir(C2H4)2,6 TpIr(C2H4)2,10 and [IrCl(coe)(CO)]231 (coe ) cyclooctene) were obtained by published procedures. TpMe2Ir(C2H4)(PMe3) (3a*). The bis(ethylene) complex 1* (0.2 g, 0.37 mmol) was dissolved in THF (20 mL), and PMe3 was added (0.40 mL, 1 M solution in THF). The mixture was stirred for 2 h at ambient temperature, and the solvent was evaporated to dryness. The oily residue was treated with petroleum ether (10 mL), and the volatiles were removed in vacuo. The resulting powder was then redissolved in acetone (10 mL), and the solution was filtered. Concentration and cooling at -20 °C afforded the product as a white-cream-colored powder in ca. 50-70% yield. 1H NMR (300 MHz, C6D6, 298 K): δ 5.78 (s, 2 H, 2 CHpyr), 5.10 (s, 1 H, CHpyr), 2.68 (m, 2 H, 2 CHolef), 2.43 (s, 6 H, 2 Me), 2.35 (s, 6 H, 2 Me), 2.22 (s, 3 H, Me), 2.04 (s, 3 H, Me), 1.56 (m, 3JPH ) 5.9 Hz, 2 H, 2 CHolef), 0.76 (d, 2JPH ) 9.6 Hz, 9 H, PMe3). 1H{31P} NMR: the C2H4 protons appear as an AA′BB′ spin system with δA 2.68 (pseudoquartet, Japp ) 4 Hz) and δB 1.56 (pseudoquartet). 31P{1H} NMR (88 MHz, C6D6, 298 K): δ -45.6 (s). 13C{1H} (50 MHz, C D , 298 K): δ 151.4 (2 CMe), 151.1 (d, 3J 6 6 PC ) 6 Hz, CMe), 144.2 (2 CMe), 141.9 (CMe), 107.7 (d, 4JPC ) 4 Hz, CHpyr), 105.7 (2 CHpyr), 17.0, 13.5, 13.1, 11.7 (s, 2:2:1:1 ratio, CMe), 12.7 (d, 1JPC ) 10 Hz, PMe3), -8.1 (d, 2JPC ) 4 Hz, 1JCH ) 145 Hz, C2H4). Anal. Calcd for C20H35N6BPIr: C, 42.5; H, 6.3; N, 13.5. Found: C, 41.9; H, 6.7; N, 13.2. TpMe2Ir(C2H4)(PMe2Ph) (3b*). This compound was obtained as pale yellow crystals from acetone in approximately the same yield by following a procedure similar to that described for complex 3a*. 1H NMR (300 MHz, C6D6, 298 K): δ 6.8 (m, 5 H, C6H5), 5.63, 5.09 (s, 2 H, 1 H, 2 CHpyr), 2.78, 1.73 (m, 2 H, 2 H, C2H4), 2.38, 2.25, 2.17, 2.03 (s, 2:1:2:1 ratio, 6 Me), 1.09 (d, 1JPH ) 9.2 Hz, 6 H, 2 PMe). 31P{1H} NMR (132 MHz, C D , 298 K): δ -40.7 (s). 13C{1H} NMR 6 6 (75 MHz, C6D6, 298 K): δ 152.1 (2 CMe), 151.3 (d, 3JPC ) 4 Hz, CMe), 144.4 (2 CMe), 142.2 (CMe), 137.7 (d, 1JPC ) 52 Hz, CarP), 130-127 (CH, Ph), 108.1 (d, 4JPC ) 4 Hz, CHpyr), 106.0 (2 CHpyr), 16.5, 13.4, 12.8, 11.3 (2:1:2:1 ratio, CMe), 11.9 (d, 1JPC ) 16 Hz, 2 PMe), -7.2 (d, 2JPC ) 4 Hz, C2H4). Anal. Calcd for C25H37N6BPIr: C, 45.8; H, 5.6; N, 12.8. Found: C, 45.9; H, 5.7; N, 12.9. TpMe2Ir(C2H4)(PEt3) (3c*). By the same general method, complex 3c* was obtained in ca. 70% yield, from Et2O, as pale yellow needles. 1H NMR (300 MHz, CDCl , 298 K): δ 5.83 (s, 2 H, 2 CH ), 5.30 (s, 3 pyr 1 H, CHpyr), 2.44 (s, 6 H, 2 Me), 2.42 (s, 6 H, 2 Me), 2.22 (m, 2 H, 2 CHolef), 2.16 (s, 3 H, Me), 2.14 (s, 3 H, Me), 1.57 (m, 2 H, 2 CHolef), 1.54 (m, 6 H, 3 PCH2CH3), 0.67 (m, 9 H, 3 PCH2CH3). 31P{1H} NMR (132 MHz, CDCl3, 298 K): δ -31.9 (s). 13C{1H} NMR (75 MHz, CDCl3, 298 K): δ 151.5, 150.7 (2:1 ratio, CMe), 144.4, 142.0 (2:1 ratio, CMe), 107.5 (d, 4JPC ) 4 Hz, CHpyr), 105.5 (2 CHpyr), 16.9, 13.3, 12.7, 11,7 (2:1:2:1 ratio, CMe), 14.9 (d, 1JPC ) 34 Hz, PCH2), 6.9 (d, 2J 2 1 PC ) 4 Hz, PCH2CH3), -10.9 (d, JPC ) 4 Hz, JCH ) 145 Hz, C2H4). Anal. Calcd for C23H41N6BPIr: C, 43.4; H, 6.5; N, 13.2. Found: C, 43.5; H, 6.6; N, 13.3. (31) Shaw, B. L.; Singleton, E. J. Chem. Soc. A 1967, 1683.

Gutie´rrez-Puebla et al. [TpMe2Ir(C2H4)]2(dmpe) (3d*). From complex 1* and dmpe, complex 3d* was obtained in ca. 45% yield in the form of small yellow needles (Et2O-CH2Cl2; -20 °C). 1H NMR (500 MHz, CDCl3, 298 K): δ 5.77 (s, 2 H, 2 CHpyr), 5.27 (s, 1 H, CHpyr), 2.34 (s, 6 H, 2 Me), 2.20 (s, 6 H, 2 Me), 2.09 (s, 3 H, Me), 2.07 (m, 2 H, 2 CHolef), 2.06 (s, 3 H, Me), 1.00 (m, 2 H, 2 CHolef), 0.66 (pseudot, JPHapp ) 4.4 Hz, 6 H, 2 PMe), 0.10 (d, JPH ) 2.3 Hz, 2 H, PCH2). 31P{1H} NMR (220 MHz, CDCl3, 298 K): δ -38.7 (s). 13C{1H} NMR (125 MHz, CDCl3, 298 K): δ 151.2, 150.9 (2:1 ratio, CMe), 143.9, 142.9 (2:1 ratio, CMe), 107.5, 105.9 (1:2 ratio, CHpyr), 20.9 (pseudot, JPCapp ) 18 Hz, PCH2), 16.8, 13.3, 12.5, 11.7 (2:1:2:1 ratio, CMe), 10.9 (pseudot, JPCapp ) 18 Hz, PMe), -9.1 (1JCH ) 146 Hz, C2H4). TpIr(C2H4)(PMe2Ph) (3b). According to the same general procedure, complex 3b was obtained in ca. 85% yield as yellow crystals from acetone. 1H NMR (500 MHz, C6D6, 296 K): δ 7.8-5.7 (m, 5 H, C6H5, m, 2:1 pattern, 9 H, CHpyr), 1.7-1.62 (m, 4 H, C2H4), 1.26 (d, 2JPH ) 9.3 Hz, 6 H, 2 PMe). 31P{1H} NMR (220 MHz, C6D6, 296 K): δ -29.0 (s). 13C{1H} NMR (120 MHz, C6D6, 296 K): δ 143104 (C6H5 and CHpyr,), 13.5 (d, 1JPC ) 37 Hz, PMe), 0.5 (d, 2JPC ) 3 Hz, C2H4). Anal. Calcd for C19H25N6BPIr: C, 39.9; H, 4.4; N, 14.7. Found: C, 39.5; H, 4.3; N, 13.9. TpIr(C2H4)(PEt3) (3c). The product was obtained as pale yellow crystals from acetone (yield 50%). 1H NMR (300 MHz, CDCl3, 296 K): 7.9-6.0 (m, 9 H, 2:1 pattern, CHpyr), 1.65 (m, 6 H, 3 PCH2), 1.18 (m, 2 H, 2 CHolef), 1.00 (m, 2 H, 2 CHolef), 0.83 (m, 9 H, 3 PCH2CH3). 31P{1H} NMR (132 MHz, CDCl , 296 K): δ -16.4 (s). 13C{1H} NMR 3 (132 MHz, CDCl3, 296 K): δ 144-104 (2:1 pattern, CHpyr,), 15.3 (d, 1J PC ) 34 Hz, PCH2), 7.8 (PCH2CH3), -2.8 (C2H4). Anal. Calcd for C17H29N6BPIr: C, 37.0; H, 5.3; N, 15.2. Found: C, 37.0; H, 5.3; N, 15.2. TpMe2Ir(C2H4)(CO) (4*). [IrCl(coe)(CO)]2 (0.3 g, 0.41 mmol) was suspended in 30 mL of THF at 0 °C. Ethylene was bubbled through the mixture for 10 min to give a colorless solution to which KTp* (0.28 g, 0.82 mmol) was added. The reaction mixture became orange and gradually evolved to a final pale reddish color after 4 h of stirring at room temperature. Volatiles were pumped off under vacuum, and the residue was extracted with 30 mL of a 1:1 mixture of Et2O and CH2Cl2. The resulting suspension was filtered through Celite to eliminate the potassium chloride, and the solution was partially evaporated until cloudiness. Cooling at -20 °C afforded 4* as a white material in 70% yield. IR (Nujol): ν(CO) 1990 cm-1. 1H NMR (500 MHz, CDCl3, 298 K): δ 5.85 (s, 2 H, 2 CHpyr), 5.50 (s, 1 H, CHpyr), 2.37 (s, 12 H, 4 Me), 2.32 (pseudoquartet, Japp ) 4.5 Hz, 2 H, 2 CHolef), 2.25 (s, 3 H, Me), 2.24 (s, 3 H, Me), 1.76 (pseudoquartet, 2 H, 2 CHolef). The AA′BB′ spin system of the C2H4 ligand has been successfully simulated: JAB ) 9.2, JAA′ ) JBB′ ) -7.4, JA′B ) JAB′ ) -4.3 Hz. 13C{1H} NMR (75 MHz, CDCl , 298 K): δ 165.3 (CO), 152.4, 150.2 3 (1:2 ratio, CMe), 143.6, 143.2 (1:2 ratio, CMe), 109.4, 105.5 (1:2 ratio, CHpyr), 15.3, 13.3, 12.4, 12.2 (2:1:1:2 ratio, CMe), 0.6 (C2H4). TpMe2IrH(CO)(COOH) (5*). Through a solution of complex 4* (0.1 g, 0.2 mmol) in THF (20 mL) was bubbled carbon monoxide for ca. 10 min. The solvent was evaporated and the residue taken up in CH2Cl2. After centrifugation, the dicloromethane was evaporated to give a white residue of spectroscopically pure 5* (yield 70%). IR (Nujol): ν(Ir-H) 2170; ν(CO) 2040; ν(COOH) 1630 cm-1. 1H NMR (500 MHz, CDCl3, 296 K): δ 8.85 (br, 1 H, COOH), 5.78, 5.72, 5.69 (s, 1 H, 1 H, 1 H, CHpyr), 2.28, 2.24, 2.23, 2.16, 2.15, 2.12 (s, 3 H each, 6 Me), -15.81 (s, 1 H, Ir-H). 13C{1H} NMR (125 MHz, CDCl3, 296 K): δ 168.3 (CO), 166.0 (2JCH ) 7 Hz, Ir-COOH), 151.5, 151.0, 150.8, 144.4, 144.3, 144.1 (CMe), 106.6, 106.5, 105.9 (CHpyr), 16.0, 15.4, 14.5, 12.7, 12.3, 12.3 (6 Me). Anal. Calcd for C17H24N6BO3Ir: C, 36.2; H, 4.3; N, 14.9. Found: C, 35.2; H, 4.2; N, 14.0. TpMe2IrH2(PMe3) (6a*). Complex 3a* (0.12 g, 0.2 mmol) was dissolved in THF (10 mL) and transferred to a Fisher-Porter bottle. The solution was pressurized with 2 atm of H2. After 2 h of stirring at room temperature, excess H2 was vented and replaced by an atmosphere of N2. The volatiles were removed in vacuo, and an 1H NMR spectrum of the residue revealed a quantitative conversion to the dihydride. A crystalline solid was obtained in 70% yield by cooling a concentrated Et2O solution at -20 °C. IR (Nujol): ν(Ir-H) 2150, 2135 cm-1. 1H NMR (300 MHz, C6D6, 298 K): δ 5.76 (s, 2 H, 2

Reactions of Tp′Ir(C2H4)(L) Complexes CHpyr), 5.47 (s, 1 H, CHpyr), 2.40, 2.30, 2.27, 2.10 (s, 1:2:2:1 ratio, 6 Me), 1.40 (d, 2JPH ) 9.6 Hz, 9 H, PMe3), -21.21 (d, 2JPH ) 26.1 Hz, 2 H, Ir-H). 31P{1H} NMR (132 MHz, C6D6, 298 K): δ -53.2 (s). 13C{1H} NMR (75 MHz, C D , 298 K): δ 150.2, 149.5 (s, d, 2:1 ratio, 6 6 3J PC ) 4 Hz, CMe), 143.7, 142.0 (s, s, 2:1 ratio, 2 CMe), 105.6 (2 CHpyr), 104, 6 (d, 4JPC ) 3 Hz, CHpyr), 23.2 (d, 1JPC ) 37 Hz, PMe3), 17.3, 17.2, 12.6, 12.3 (s, 1:2:2:1 ratio, CMe). Anal. Calcd for C18H33N6BPIr: C, 38.1; H, 5.8; N, 14.8. Found: C, 38.4; H, 6.0; N, 14.4. TpMe2IrH2(PMe2Ph) (6b*). This complex was obtained by following the method described for the previous dihydride. Colorless crystals (70% yield) were obtained from Et2O. IR (Nujol): ν(Ir-H) 2160, 2145 cm-1. 1H NMR (500 MHz, C6D6, 298 K): δ 7.5-6.8 (m, 5 H, C6H5), 5.66 (s, 2 H, 2 CHpyr), 5.53 (s, 1 H, CHpyr), 2.47, 2.31, 2.14, 1.95 (s, 1:2:1:2 ratio, 6 Me), 1.88 (d, 2JPH ) 9.0 Hz, 6 H, 2 PMe), -21.16 (d, 2J 31P{1H} NMR (220 MHz, C D , 298 K): PH ) 26.8 Hz, 2 H, Ir-H). 6 6 δ -38.3 (s). 13C{1H} NMR (120 MHz, C6D6, 298 K): δ 150.6, 149.7 (2:1 ratio, CMe), 143.5, 142.0 (2:1 ratio, CMe), 139.3 (d, 1JPC ) 45 Hz, CarP), 131-127 (CH, Ph), 105.7, 104.8 (2:1 ratio, CHpyr), 25.7 (d, 1J PC ) 41 Hz, PMe), 17.5, 16.6, 12.7, 12.2 (1:2:2:1 ratio, CMe). Anal. Calcd for C23H35N6BPIr: C, 43.9; H, 5.6; N, 13.3. Found: C, 44.1; H, 5.8; N, 13.3. [TpMe2IrH2]2(dmpe) (6d*). According to the already described general method, this complex was obtained in 70% yield as white crystals from Et2O-CH2Cl2. IR (Nujol): ν(Ir-H) 2165, 2130, 2110 cm-1. 1H NMR (500 MHz, CDCl3, 298 K): δ 5.75 (s, 2 H, 2 CHpyr), 5.57 (s, 1 H, CHpyr), 2.37 (s, 6 H, 2 Me), 2.19 (s, 3 H, Me), 2.10 (s, 6 H, 2 Me), 2.07 (s, 3 H, Me), 1.25 (pseudot, JPHapp ) 3.7 Hz, 6 H, 2 PMe), 0.76 (s, 2 H, PCH2), -21.87 (filled-in d, JPHapp ) 24.5 Hz, 2 H, Ir-H). 31P{1H} NMR (220 MHz, CDCl3, 298 K): δ -40.0 (s). 13C{1H} NMR (125 MHz, CDCl3, 298 K): δ 150.2, 149.2 (2:1 ratio, CMe), 143.8, 142.6 (2:1 ratio, CMe), 105.5, 104.1 (2:1 ratio, CHpyr), 25.7 (pseudot, JPCapp ) 16 Hz, PCH2), 22.7 (pseudot, JPCapp ) 19 Hz, PMe), 17.2, 17.0, 12.7, 12.6 (1:2:1:2 ratio, CMe). Anal. Calcd for C36H64N12B2P2Ir2: C, 38.2; H, 5.6; N, 14.8. Found: C, 38.6; H, 5.6; N, 14.3. TpIrH2(PMe2Ph) (6b). The above procedure gave crude white material with a quantitative yield. Colorless crystals may be obtained from Et2O at -20 °C. IR (Nujol): ν(Ir-H) 2138 cm-1. 1H NMR (500 MHz, CDCl3, 298 K): δ 8.0-7.0 (m, 11 H, C6H5 and 3,5-CHpyr), 5.84 (m, 2 H, 4-CHpyr), 5.70 (m, 1 H, 4-CHpyr), 1.69 (d, 2JPH ) 9.8 Hz, 6 H, 2 PMe), -20.34 (d, 2JPH ) 25.2 Hz, 2 H, Ir-H). 31P {1H} NMR (220 MHz, CDCl3, 298 K): δ -31.0 (s). 13C{1H} NMR (125 MHz, CDCl3, 298 K): δ 146.1, 143.5 (d, s, 3JPC ) 3 Hz, 1:2 ratio, 3-CHpyr), 138.2 (d, 1JPC ) 51 Hz, CarP), 134.0, 133.1 (s, s, 2:1 ratio, 5-CHpyr), 131-127 (C6H5), 105.4, 104.9 (s, s, 1:2 ratio, 4-CHpyr), 21.0 (d, 1JPC ) 40 Hz, PMe). Anal. Calcd for C17H23N6BPIr‚1/2Et2O: C, 39.2; H, 4.8; N, 14.4. Found: C, 40.0; H, 5.3; N, 13.9. TpMe2IrH2(CO) (7*). This complex was obtained similarly to the above dihydrides. The yield was quantitative by NMR. IR (Nujol): ν(CO) 2008; ν(Ir-H) 2150, 2120 cm-1. 1H NMR (500 MHz, C6D6, 298 K): δ 5.54 (s, 2 H, 2 CHpyr), 5.42 (s, 1 H, CHpyr), 2.25, 2.24, 2.14, 2.02 (s, 2:1:2:1 ratio, 6 Me), -16.50 (s, 2 H, Ir-H). 13C{1H} NMR (125 MHz, C6D6, 298 K): δ 170.0 (CO), 151.0, 150.4 (1:2 ratio, CMe), 143.4, 143.3 (1:2 ratio, CHpyr), 105.5, 105.4 (2:1 ratio, CHpyr), 17.0, 15.3, 12.0, 11.9 (1:2:1:2 ratio, CMe). TpMe2IrHCl(PMe3) (8a*). A 0.02 g sample of complex 6a* was dissolved in a mixture of CDCl3 (0.5 mL) and CCl4 (0.2 mL), and the resulting solution was transferred to an NMR tube. Heating at 80-90 °C, with continuous monitoring of the reaction by 1H NMR spectroscopy, afforded the title compound. The solvent was evaporated to dryness to give the product as a white powder. IR (Nujol): ν(Ir-H) 2200 cm-1. 1H NMR (500 MHz, CDCl3, 298 K): δ 5.79, 5,72, 5,65 (s, 1 H, 1H, 1 H, CHpyr), 2.58, 2.43, 2.41, 2.37, 2.23, 2,19 (s, 3 H each, 6 Me), 1.55 (d, 2JPH ) 9.5 Hz, 9 H, PMe3), -23.43 (d, 2JPH ) 21.8 Hz, 2 H, Ir-H). 31P{1H} NMR (220 MHz, CDCl3, 298 K): δ -50.8 (s). 13C{1H} NMR (125 MHz, CDCl3, 298 K): δ 152.4, 151.6, 151.1, 145.1, 143.4, 142.6 (CMe), 107.8, 106.7, 106.0 (CHpyr), 18.1, 17.8, 17.5, 13.1, 12.6, 12.4 (CMe), 14.9 (d, 1JPC ) 46 Hz, PMe3). Anal. Calcd for C18H32N6BClPIr: C, 35.9; H, 5.3; N, 14.0. Found: C, 35.7; H, 5.5; N, 11.6.

Inorganic Chemistry, Vol. 37, No. 18, 1998 4545 TpMe2IrHCl(PMe2Ph) (8b*). This compound was obtained in the same fashion as the monochloride described above. The crude sample can be crystallized from Et2O at -20 °C. IR (Nujol): ν(Ir-H) 2225 cm-1. 1H NMR (500 MHz, C6D6, 298 K): δ 7.3-6.9 (m, 5 H, C6H5), 5.71, 5.48, 5.24 (s, 1 H, 1 H, 1 H, CHpyr) 2.7-1.4 (s, 3 H each, 6 Me), 1.96 (d, 2JPH ) 10.1 Hz, 3 H, PMe), 1.44 (d, 2JPH ) 10.0 Hz, 3 H, PMe), -22.93 (d, 2JPH ) 21.3 Hz, 1 H, Ir-H). 31P{1H} NMR (220 MHz, C6D6, 298 K): δ -39.0 (s). 13C{1H} NMR (125 MHz, C6D6, 298 K): δ 153.1, 151.8, 151.6, 143.9, 143.3, 142.1 (s, s, d, s, s, s, 3JPC ) 3 Hz, CMe), 137.6 (d, 1JPC ) 52 Hz, CarP), 130-127 (CH, Ph), 108.0, 106.3, 106.2 (CHpyr), 18.5 (d, 1JPC ) 42 Hz, PMe), 16.0 (d, 1JPC ) 37 Hz, PMe), 16.2, 15.1, 15.1, 12.8, 12.6, 12.2 (CMe). Anal. Calcd for C23H34N6BClPIr: C, 41.6; H, 5.1; N, 12.7. Found: C, 42.0; H, 5.2; N, 12.3. TpMe2IrH2(C2H4) (9*). A solution of complex 1* (0.3 g, 0.55 mmol) in THF (30 mL) was pressurized with H2 (2 atm). After 1 h of stirring at room temperature, the H2 atmosphere was replaced by N2, and the solvent was evaporated. 1H NMR analysis of the crude material was in accord with its being a 1:1.5 mixture of complexes 9* and 10*. Column cromatography (silica gel as support; petroleum ether as eluent) allowed the separation of the two compounds with the ethyl derivative being the first to drain down the column. The individual compounds were obtained as colorless crystals by crystallization from petroleum ether. IR (Nujol): ν(Ir-H) 2190, 2170 cm-1. 1H NMR (200 MHz, C6D6, 298 K): δ 5.61 (s, 2 H, 2 CHpyr), 5.48 (s, 1 H, CHpyr), 3.46 (s, 4 H, C2H4), 2.35 (s, 3 H, Me), 2.23 (s, 6 H, 2 Me), 2.09 (s, 3 H, Me), 1.92 (s, 6 H, 2 Me), -20.77 (s, 2 H, Ir-H).13C{1H} NMR (50 MHz, C6D6, 298 K): δ 150.6, 150.2 (2:1 ratio, CMe), 143.2, 142.8 (2:1 ratio, CMe), 106.1, 105.2 (2:1 ratio, CHpyr), 35.6 (C2H4), 17.5, 13.9, 12.3, 12.2 (1:2:1:2 ratio, CMe). Anal. Calcd for C17H28N6BIr: C, 39.2; H, 5.4; N, 16.1. Found: C, 39.5; H, 5.6; N, 16.3. TpMe2IrH(CH2CH3)(C2H4) (10*). IR (Nujol): ν(Ir-H) 2195 cm-1. 1H NMR (200 MHz, C D , 298 K): 5,74, 5.62, 5.39 (s, 1 H, 1 H, 1 H, 6 6 CHpyr), 3.65 (m, AA′ part of an AA′XX′ spin system, 2 H, 2 CHolef), 2.80 (m, XX′ part of an AA′XX′ spin system, 2 H, 2 CHolef), 2.49, 2.30, 2.22, 2.13, 2.08, 1.72 (s, 3 H each, 6 Me), 2.43 (dq, 2JAB ) 9.8, 3J HH ) 7.0 Hz, 1 H, CHAHBCH3), 0.68 (t, 3 H, CH2CH3), 0.50 (dq, 1 H, CHACHBCH3), -17.52 (s, 1 H, IrH). 13C{1H} NMR (50 MHz, THFd8, 298 K): δ 155-140 (6 CMe), 109.0, 107.7, 107.4 (CHpyr), 43.6 (C2H4), 16.4 (IrCH2CH3), 15.4, 15.0, 14.3, 13.4, 13.1 (1:1:1:2 ratio, CMe), -17.9 (IrCH2CH3). Anal. Calcd for C19H32N6BIr: C, 41.7; H, 5.8; N, 15.3. Found: C, 41.8; H, 6.0; N, 14.9. TpIrH(CH2CH3)(C2H4) (10). Under the same experimental conditions described above, complex 1 was hydrogenated to give complex 10 as the sole product. Pale yellow crystals were obtained from concentrated solutions in petroleum ether-Et2O (2:1). IR (Nujol): ν(Ir-H) 2200 cm-1. 1H NMR (200 MHz, C6D6, 298 K): δ 7.90, 7.50, 7.38, 7.32, 7.30, 6.80 (d, 3JHH ) 2.5 Hz, 1 H each, 3,5-CHpyr), 5.9, 5.76, 5.74 (t, 1 H each, 4-CHpyr), 3.0 (m, 4 H, C2H4), 1.41, 1.17 (m, m, 1 H each, CH2CH3), 1.23 (t, 3JHH ) 7.0 Hz, 3 H, CH2CH3), -16.2 (s, 1 H, IrH). 13C{1H} NMR (50 MHz, C6D6, 298 K): δ 142.7, 140.8, 137.8, 134.1, 105.9, 105.8, 105.2 (CHpyr), 43.6 (C2H4), 19.7 (IrCH2CH3), -13.9 (IrCH2CH3). TpMe2IrH(CH2CH3)(PMe3) (11*). Complex 9* (0.3 g, 0.55 mmol) was dissolved in neat PMe3 (1 mL), and the resulting mixture was heated at 60 °C for 8 h (sealed ampule). The volatiles were removed in vacuo, and the residue was extracted with petroleum ether. Concentration and cooling at -20 °C afforded white crystals in 70% yield. IR (Nujol): ν(Ir-H) 2170 cm-1. 1H NMR (300 MHz, C6D6, 298 K): δ 5.75, 5.63, 5.57 (s, 1 H, 1 H, 1 H, CHpyr), 2.58 (m, 2 H, 2 IrCH2), 2.53, 2.38, 2.27, 2.20, 2.14, 2.10 (s, 3 H each, 6 Me), 1.45 (t, 3J 2 HH ) 7.5 Hz, 3 H, IrCH2CH3), 1.34 (d, JPH ) 9.1 Hz, 9 H, PMe3), -23.19 (d, 2JPH ) 25.9 Hz, 1 H, IrH). 31P{1H} NMR (132 MHz, C6D6, 298 K): δ -51.6 (s). 13C{1H} NMR (75 MHz, C6D6, 298 K): δ 107.1, 105.8 (1:2 ratio, CHpyr), 21.9 (IrCH2CH3), 19.5 (d, 1JPC ) 36 Hz, PMe3), 17.0, 15.2, 15.0, 12.8, 12.7, 12.5 (CMe), -24.2 (d, 2JPC ) 6 Hz, IrCH2CH3). TpMe2Ir(CH2CH3)2(PMe3) (12*). According to an analogous procedure, but starting with 10* (0.1 g, 0.18 mmol), white crystals of the bis(ethyl) complex were obtained from petroleum ether (90% yield). 1H NMR (500 MHz, C D , 298 K): δ 5.68 (s, 2 H, 2 CH ), 5.61 (s, 6 6 pyr

4546 Inorganic Chemistry, Vol. 37, No. 18, 1998 1 H, CHpyr), 2.62 (s, 3 H, Me), 2.51 (dq, 2JAB ) 12.6, 3JHH ) 7.5 Hz, 2 H, 2 CHAHBCH3), 2.35 (dquint, 3JPH ) 7.5 Hz, 2 H, 2 CHAHBCH3), 2.26 (s, 6 H, 2 Me), 2.18 (s, 6 H, 2 Me), 2.09 (s, 3 H, Me), 1.17 (d, 2 JPH ) 8.7 Hz, 9 H, PMe3), 0.92 (t, 6 H, 2 CH2CH3). 31P{1H} NMR (88 MHz, C6D6, 298 K): δ -49.8 (s). 13C{1H} NMR (125 MHz, C6D6, 298 K): δ 150.2, 148.9 (d, s, 1:2 ratio, 3JPC ) 4 Hz, CMe), 142.8, 142.1, (2:1 ratio, Me), 107.6, 107.4 (d, s, 1:2 ratio, 4JPC ) 3 Hz), 16.3 (d, 3JPC ) 1 Hz, IrCH2CH3), 15.7, 13.7, 12.9, 12.8 (2:1:2:1 ratio, CMe), 15.6 (d, 1JPC ) 36 Hz, PMe3), -19.5 (d, 2JPC ) 7 Hz, IrCH2CH3). Anal. Calcd for C22H41N6BPIr: C, 42.3; H, 6.6; N, 13.5. Found: C, 42.5; H, 6.7; N, 13.5. TpMe2IrH(CHdCH2)(PMe3) (13a*). Complex 3a* was dissolved in 0.5 mL of C6D6 and the resulting solution transferred to a NMR tube. Heating at 60 °C was monitored periodically until the transformation was completed. The solvent was removed in vacuo to give complex 13a* in quantitative yield. This white material was pure enough for most purposes. It can be recrystallized from acetone at -20 °C but with appreciable losses. IR (Nujol): ν(Ir-H) 2170 cm-1. 1H NMR (300 MHz, C D , 298 K): δ 8.51 (ddd, 3J 3 6 6 AX ) 18.2, JAM ) 10.6, 3JPA ) 2.0 Hz, 1 H, HA), 6.42 (dd, 2JMX ) 3.8 Hz, 1 H, HM), 5,78, 5.65, 5.50 (s, 1 H, 1 H, 1 H, CHpyr), 5.52 (dd, 1 H, HX), 2.56, 2.38, 2.30, 2.23, 2.16, 2.08 (s, 3 H each, 6 Me), 1.33 (d, 2JPH ) 9.4 Hz, 9 H, PMe3), -21.60 (d, 2JPH ) 25.7 Hz, 1 H, Ir-H). 31P{1H} NMR (132 MHz, C6D6, 298 K): δ -49.7 (s). 13C{1H} NMR (75 MHz, C6D6, 298 K): δ 150.7, 150.4, 150.1, 143.9, 143.4, 142.1 (CMe), 130.3 (d, 2JPC ) 10 Hz, IrCHdCH2), 121.1 (d, 3JPC ) 3 Hz, IrCH)CH2), 107.0, 106.0, 105.9 (3 CHpyr), 18.9 (d, 1JPC ) 37 Hz, PMe3), 16.9, 16.2, 15.7, 12.7, 12.6, 12.4 (6 CMe).

TpMe2IrH(CHdCH2)(PMe2Ph) (13b*). In the same way as described above for the PMe3 analogue, complex 13b* was obtained in quantitative yield. The analytical sample, white crystals, was recrystallized from acetone. IR (Nujol): ν(Ir-H) 2195 cm-1. 1H NMR (500 MHz, C6D6, 298 K): δ 8.61 (ddd, 3JAX ) 18.2, 3JAM ) 10.6, 3JPA ) 3.3 Hz, 1 H, HA), 7.4-6.8 (m, 5 H, C6H5), 6.39 (dd, 2JMX ) 3.6 Hz, 1 H, HM), 5,80, 5.56, 5.41 (s, 1 H, 1 H, 1 H, CHpyr), 5.55 (dd, 1 H, HX), 2.60, 2.34, 2.33, 2.23, 2.11, 1.63 (s, 3 H each, 6 Me), 1.98, 1.52 (d, 2J 2 PH ) 9.3 Hz, 3 H each, 2 PMe), -21.38 (d, JPH ) 25.1 Hz, 1 H, IrH). 31P{1H} NMR (132 MHz, C6D6, 298 K): δ -36.1 (s). 13C{1H} NMR (75 MHz, C6D6, 298 K): δ 150.8, 150.6, 150.5, 143.4, 142.1 (CMe), 138.5 (d, 1JPC ) 49 Hz, CarP), 130-127 (CH, Ph), 130.3 (d, 2J PC ) 10 Hz, IrCHdCH2), 121.0 (IrCHdCH2), 107.1, 106.0, 105.8 (3 CHpyr), 21.8 (d, 1JPC ) 42 Hz, PMe), 16.9 (d, 1JPC ) 36 Hz, PMe), 16.5, 15.9, 15.8, 12.8, 12.7, 12.4 (CMe). Anal. Calcd for C25H37N6BPIr: C, 45.8; H, 5.7; N, 12.8. Found: C, 45.5; H, 5.8; N, 12.2. TpMe2IrH(CHdCH2)(CO) (14*). Complex 4* was dissolved in C6H12, and the mixture was heated at 120 °C (sealed ampule) until complete disappearance of the starting material (NMR monitoring). The hydride-vinyl complex was obtained in ca. 70% yield along with some unidentified material. IR (petroleum ether): ν(CO) 2020 cm-1. 1H NMR (500 MHz, CDCl3, 298 K): δ 7.14 (dd, 3JAX ) 18.1, 3JAM ) 10.4 Hz, 1 H, HA), 5.98 (dd, 3JMX ) 2.7 Hz, 1 H, HM), 5.81, 5.78, 5.77

Gutie´rrez-Puebla et al. (s, 1 H, 1 H, 1 H, CHpyr), 5.37 (dd, 1 H, HX), 2.4-2.2 (6 s, 3 H each, 6 Me), -16.62 (s, 1 H, IrH). 13C{1H} NMR (125 MHz, CDCl3, 298 K): δ 169.1 (CO), 151.3, 151.1, 150.4, 143.8, 143.7, 143.4 (CMe), 123.7, 123.6 (IrCHdCH2), 106.4, 106.3, 105.5 (CHpyr), 16-12 (6 Me). X-ray Structure Determination of 3b*. A summary of the fundamental crystal data is given in Table 1. A yellow crystal of prismatic shape was coated with an epoxy resin and mounted in a κ diffractometer. The cell dimensions were refined by least-squares fitting of the values of 25 reflections. The intensities were corrected for Lorentz and polarization effects. Scattering factors for neutral atoms and anomalous dispersion corrections for Ir and P were taken from ref 32. The structure was solved by Patterson and Fourier methods. The compound crystallizes with 1/4 of CH3COCH3 per formula unit. An empirical absorption correction33 was applied at the end of the isotropic refinement. Some nonresolvable disorder from the thermal motion was found around the CH3COCH3 molecule of crystallization, and because of this, the atoms of this molecule were refined only isotropically. No trend in ∆F vs Fo or (sin θ)/λ was observed. Final refinement with fixed isotropic temperature factors and coordinates for hydrogen atoms gave R ) 0.43. Final difference calculations were carried out with the X-ray 80 system.34 X-ray Structure Determination of 13b*. A summary of the fundamental crystal data is given in Table 1. A colorless crystal of prismatic shape was coated with an epoxy resin and mounted in a κ diffractometer. The cell dimensions were refined by least-squares fitting of the values of 25 reflections with a 2θ range of 12-28°. The intensities were corrected for Lorentz and polarization effects. Scattering factors for neutral atoms and anomalous dispersion corrections for Ir and P were taken from ref 32. The structure was solved by Patterson and Fourier methods. An empirical absorption correction33 was applied at the end of the isotropic refinement. To prevent bias on ∆F vs Fo or (sin θ)/λ, weights were assigned as w ) 1/(a + bFo)2, with the following coefficients: for Fo < 45, a ) 10.2 and b ) -0.22; for Fo > 45, a ) 1.05 and b ) 0.01. A final mixed refinement was undertaken. Hydrogen atoms were included with fixed isotropic contributions at their calculated positions, except the H1 atom, which was located in a difference Fourier map and whose coordinates were refined. Final difference synthesis showed no significant electron density. Most of the calculations were carried out with the X-ray 80 system.34

Acknowledgment. This research was supported by DGES Projects PB93-0921 and PB94-1445. M.C.N., P.J.P., and L.R. thank the Ministerio de Educacio´n y Cultura for research fellowships. We also acknowledge the University of Sevilla for free access to its analytical and NMR facilities. Supporting Information Available: For 3b* and 13b*, tables of positional and thermal parameters, fractional coordinates, and bond lengths and angles (20 pages). Ordering information is given on any current masthead page. IC9800785 (32) Cromer, D. T.; Waber, J. T. International Tables for X-ray Crystallography; Kynoch Press: Birmingham, U.K., 1974; Vol. IV, p 72. (33) Walker, N.; Stuart, D. Acta Crystallogr. 1983, A39, 158. (34) Stewart, J. M. The X-ray 80 system; Computer Science Center, University of Maryland: College Park, MD, 1985.